Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Draft version July 5, 2018

Typeset using LATEX twocolumn style in AASTeX62

Extremely Large Extreme-ultraviolet Late Phase Powered by Intense Early Heating in a Non-eruptive Solar Flare
Yu Dai,1, 2, 3 Mingde Ding,1, 2 Weiguo Zong,3 and Kai E. Yang1, 2
1 School of Astronomy and Space Science, Nanjing University, Nanjing 210023, People’s Republic of China
2 Key Laboratory of Modern Astronomy and Astrophysics (Nanjing University), Ministry of Education, Nanjing 210023, People’s Republic
of China
arXiv:1807.01315v1 [astro-ph.SR] 3 Jul 2018

3 Key Laboratory of Space Weather, National Center for Space Weather, China Meteorological Administration, Beijing 100081, People’s

Republic of China

ABSTRACT
We analyze and model an M1.2 non-eruptive solar flare on 2011 September 9. The flare exhibits a
strong late-phase peak of the warm coronal emissions (∼3 MK) in extreme-ultraviolet (EUV), with peak
emission over 1.3 times that of the main flare peak. Multiple flare ribbons are observed, whose evolution
indicates a two-stage energy release process. A non-linear force-free field (NLFFF) extrapolation reveals
the existence of a magnetic null point, a fan-spine structure, and two flux ropes embedded in the fan
dome. Magnetic reconnections involved in the flare are driven by the destabilization and rise of one of
the flux ropes. In the first stage, the fast ascending flux rope drives reconnections at the null point and
the surrounding quasi-separatrix layer (QSL), while in the second stage, reconnection mainly occurs
between the two legs of the field lines stretched by the eventually stopped flux rope. The late-phase
loops are mainly produced by the first-stage QSL reconnection, while the second-stage reconnection is
responsible for the heating of main flaring loops. The first-stage reconnection is believed to be more
powerful, leading to an extremely strong EUV late phase. We find that the delayed occurrence of the
late-phase peak is mainly due to the long cooling process of the long late-phase loops. Using the model
enthalpy-based thermal evolution of loops (EBTEL), we model the EUV emissions from a late-phase
loop. The modeling reveals a peak heating rate of 1.1 erg cm−3 s−1 for the late-phase loop, which is
obviously higher than previous values.

Keywords: Sun: corona — Sun: flares — Sun: UV radiation — Sun: magnetic fields — hydrodynamics

1. INTRODUCTION hard X-ray (HXR) and chromospheric lines (e.g., He II),


Solar flares, one of the most dynamic phenomena in indicating a prompt response of the solar lower atmo-
the solar atmosphere, are manifested as rapid and sig- sphere to nonthermal electron bombardment and/or
nificant enhancements of electromagnetic radiation in a thermal conduction caused by the initial energy re-
wide wavelength range. It is widely accepted that so- lease. Since the energy transported downward cannot
lar flares are a result of rapid release of free magnetic be effectively radiated away at the solar lower atmo-
energy stored in the solar corona, and magnetic recon- sphere (Antiochos & Sturrock 1976), a hot evapora-
nection plays a crucial role in converting the magnetic tive flow is consequently driven into the flare loops
energy to plasma heating, particle acceleration, and bulk (Antiochos & Sturrock 1978), which then brighten up
mass motions (Parker 1963). in soft X-ray (SXR) and coronal extreme-ultraviolet
According to the standard two-ribbon solar flare (EUV) lines. As the flare plasma cools down, these
model, which is conventionally called the CSHKP emissions peak sequentially in an order of decreas-
model (Carmichael 1964; Sturrock 1966; Hirayama 1974; ing temperatures, and then decay gradually toward
Kopp & Pneuman 1976), the evolution of a solar flare the background levels, constituting the gradual phase
can be divided into two phases: an impulsive phase, (Chamberlin et al. 2012).
and a following gradual phase. The impulsive phase The original CSHKP model and its variants are a
is characterized by a rapid increase of the emissions in two-dimensional (2D) model in nature. Nevertheless,
a real solar flare takes place in a three-dimensional (3D)
magnetic field, making its evolution fairly more com-
Corresponding author: Yu Dai plicated than that expected based on a simplified 2D
ydai@nju.edu.cn model. By using EUV irradiance observations with the
EUV Variability Experiment (EVE; Woods et al. 2012)
2 Dai et al.

on board the recently launched Solar Dynamics Obser- tion of EUV late-phase flares. In particular, Dai & Ding
vatory (SDO ; Pesnell et al. 2012) mission, Woods et al. (2018) found that even with an equal energy partition
(2011) discovered a new phenomenon in some flares, between the late-phase loop and main flaring loop, the
namely, an “EUV late phase”. Observationally, the warm coronal late-phase peak is still significantly lower
EUV late phase is seen as a second peak in the warm than the corresponding main flare peak. This result may
coronal emissions (∼3 MK) several tens of minutes to a offer a clue to why EUV late-phase flares are rare among
few hours after the GOES SXR peak. There are, how- all solar flares. Nevertheless, as shown in the statistics
ever, no significant enhancements of the SXR or hot in Woods et al. (2011), the ratio of the late-phase peak
coronal emissions (∼ 10 MK or higher) in the EUV late to main flare peak in their sample can be as high as 4.
phase, and spatially resolved imaging observations such Liu et al. (2015) have recently reported a long-duration
as those from the Atmospheric Imaging Assembly (AIA; but non-eruptive flare in which the late-phase peak is
Lemen et al. 2012) also on board SDO reveal that the ∼ 2.1 times higher than the main flare peak. For this
secondary late-phase emission comes from another set reason they referred to it as an “extremely large EUV
of flare loops higher and longer than the main flaring late phase”. By analyzing the kinematic and thermody-
loops. namic evolution of the flare, Liu et al. (2015) proposed
In a preliminary statistics of 25 EUV late-phase solar that an “erupted-but-failed” hot structure, which is
flares occurring during the first year of SDO normal op- most presumably a magnetic flux rope, serves as a per-
erations, Woods et al. (2011) found that about half of sistent heating agent for this extremely large EUV late
them took place in two active regions (ARs), implying a phase. Wang et al. (2016) argued that a strong con-
specific magnetic configuration of the ARs in which EUV straint of the overlying arcades may prevent a flux rope
late-phase flares are preferentially produced. In case from escaping, and cause the energy carried by the flux
studies, photospheric magnetograms of the flare-hosting rope to be re-deposited into the thermal emissions that
ARs reveal that the EUV late-phase flares are commonly form a stronger late phase. This argument was fur-
involved in a multipolar magnetic field, which exhibits ther validated through a comparative study of 12 EUV
either a symmetric or asymmetric quadrupolar configu- late-phase flares (Wang et al. 2016) showing that the
ration (Hock et al. 2012; Liu et al. 2013), or a parasitic relative late-phase peaks of the non-eruptive flares are
polarity embedded in a large-scale bipolar magnetic field systematically stronger than those of the eruptive flares.
(Dai et al. 2013; Sun et al. 2013; Masson et al. 2017). In addition to the persistent heating scenario proposed
Such a magnetic configuration facilitates the existence by Liu et al. (2015), we are wondering if an impulsive
of two sets of loops that are magnetically related but heating on the late-phase loops, which is temporally
distinct in length, as further confirmed by the force-free close to the main flare heating, can also produce a signif-
coronal magnetic field extrapolations (Sun et al. 2013; icantly large EUV late phase. In this paper, we report
Li et al. 2014; Masson et al. 2017). observations of another non-eruptive solar flare that ex-
The EUV irradiance output from the Sun can drive hibits an extremely large EUV late phase. The detailed
disturbances in Earth’s ionosphere and thermosphere, data analysis shows that this large EUV late phase is
particularly during solar flares (Kane & Donnelly 1971). mainly powered by an intense heating even earlier than
Hence, the origin of EUV late phase in a solar flare is of the main flare heating. The paper is organized as fol-
interest for its potential geo-effectiveness. Since many lows. In Section 2, we describe the evolution of the flare.
theoretical works on the cooling of flare plasmas had In Section 3, we explore the magnetic topology of the
shown that the cooling time of a flare loop increases as flare-hosting AR based on a non-linear force-free field
the loop length increases (e.g., Cargill et al. 1995), it was (NLFFF) extrapolation, and in Section 4, we use the
naturally proposed that the delayed occurrence of an EBTEL model to synthesize the EUV emissions from a
EUV late phase is the result of a long-lasting cooling pro- late-phase loop. The results are discussed in Section 5,
cess in the long late-phase loops initially heated simul- and a brief summary is presented in Section 6.
taneously with the main flaring loops (Liu et al. 2013;
Masson et al. 2017). In this sense, the late phase is actu- 2. EVOLUTION OF THE FLARE
ally the gradual phase in long flare loops. Nevertheless, The flare-hosting AR NOAA 11283 has been ex-
there was also accumulating evidence for the existence of tensively studied in the literature (e.g., Feng et al.
a delayed secondary heating, which might be responsible 2013; Jiang et al. 2014; Liu et al. 2014; Xu et al. 2014;
for the production of the EUV late phase (Hock et al. Romano et al. 2015; Ruan et al. 2015; Zhang et al.
2012; Dai et al. 2013). In fact, the two mechanisms are 2015). Its passage through the visible disk from 2011
not mutually exclusive, and in some EUV late-phase August 31 to September 11 can be divided into two
flares they may both play a role (Sun et al. 2013). periods of roughly equal durations. During the first
By using the model called enthalpy-based thermal period, the AR was simply bipolar, indicative of a low
evolution of loops (EBTEL; Klimchuk et al. 2008; flare level. As revealed by the GOES flare catalog, the
Cargill et al. 2012; Barnes et al. 2016), Li et al. (2014) AR had just harbored eight flares of classes at most
and Dai & Ding (2018) numerically probed the produc- C2.2 during this period. From September 4 to 5, a
EUV LATE PHASE IN A NON-ERUPTIVE FLARE 3

GOES Flux (W m−2) (a) 1−8 Å 2.1. Extremely Large EUV Late Phase

GOES Class
10−5 M
Figure 1(a) shows the GOES 1–8 Å light curve of
10−6 C
the flare under study. Following a small bump around
12:30 UT, the SXR flux starts to rise rapidly from
11:30 12:00 12:30
40
13:00 13:30 14:00 14:30
15 ∼12:39 UT and reaches its peak at 12:49:19 UT (out-
(b) EVE 304 Å AIA 304 Å
30 He II log(T/K)=4.7
10
lined by the vertical dashed line). It is clearly seen that
20
the rise phase of SXR exhibits a two-stage evolution: a
10 5
0 very prompt rise followed by a less impulsive one. After
0
−10
30 the peak, the SXR emission also reveals two distinct de-
3 (c) EVE 133 Å AIA 131 Å
2
Fe XX/XXIII log(T/K)=7.0
20
cay periods: a rapid decay, and a following much longer
1
and less sloped one lasting until ∼13:40 UT when the
10
SXR emission is elevated again.
EVE Irradiance (µW m−2)

AIA Intensity (106 DN s−1)


0
3 8 To further study the thermal evolution of the flare,
(d) EVE 94 Å AIA 94 Å
2 Fe XVIII log(T/K)=6.8 6 we plot in Figures 1(b)–(f) the time profiles of the
1
4 background-subtracted irradiance in several EVE spec-
0
2 tral lines (we adopt a 5-minute average between 11:30
0 and 11:35 UT as the pre-flare background). EVE mea-
6 (e) EVE 335 Å AIA 335 Å 4
Fe XVI log(T/K)=6.5 3
sures full-disk integrated solar EUV irradiance from 1
4
2 to 1050 Å with 1 Å spectral resolution and 10 s time
2
0
1 cadence. By integrating EVE spectra (EVS files) over
0 specified spectral windows, the irradiance of some “iso-
(f) EVE 171 Å AIA 171 Å 40
2 Fe IX log(T/K)=5.9 30
lated” EUV lines (EVL files) can be derived. Here we
1 20 choose five lines from the Multiple EUV Grating Spec-
0
10 trographs (MEGS)-A component of EVE, which covers
0
−1 a wavelength range of 65–370 Å with a nearly 100% duty
11:30 12:00 12:30 13:00 13:30 14:00 14:30
Time (UT on 2011−Sep−09) cycle.
Several emission patterns of the flare can be seen from
Figure 1. Time profiles of the GOES 1–8 Å SXR flux (a), the EVE profiles. First, the cold chromospheric EVE
and the EVE full-disk irradiance (black) and AIA sub-region 304 Å emission (dominated by the He II λ303.8 Å line,
intensities (colored) in several lines and passbands (b)–(f) log(T /K) ∼ 4.7, Figure 1(b)) shows a very impulsive rise
for the 2011 September 9 M1.2 flare. The vertical dashed that quickly reaches its peak at 12:42:54 UT, character-
lines mark the peak times of the GOES flux and AIA in- izing the flare’s impulsive phase. In spite of a high noise
tensities. In deriving the EVE profiles, a line-dependent level in the original data, the smoothed profile (after
smoothing boxcar ranging from 60 to 300 s is applied to applying a 300 s smoothing boxcar) of the cool coronal
the original data points (plus signs) to enhance the ratio of EVE 171 Å emission (Fe IX λ171.1 Å, log(T /K) ∼ 5.9,
signal to noise. The region over which AIA pixel count rates Figure 1(f)) also reveals an obvious rise with a local
are summed covers a field-of-view (FOV) of 200′′ ×200′′ en- maximum near the peak time in EVE 304 Å. This im-
closing the flare-hosting AR. Note that the EVE and AIA
plies that the chromosphere and transition region (TR),
as responding to the initial energy release, are instantly
profiles are plotted with the corresponding background levels
heated to at least ∼ 1 MK (Chamberlin et al. 2012;
subtracted. For clarity, in each panel the background levels
Milligan et al. 2012).
for the EVE and AIA profiles are plotted (horizontal dashed
Second, the time profile of the hot coronal EVE 133 Å
lines) with a vertical offset.
emission (blended by Fe XX λ132.8 Å and Fe XXIII
λ132.9 Å, log(T /K) ∼ 7, Figure 1(c)), closely resem-
persistent emergence of positive magnetic flux inside bling the GOES SXR light curve thanks to their similar
the leading negative polarity of the AR was observed, temperature sensitivity, also exhibits a dual-rise followed
which not only complicated the magnetic topology of by a dual-decay, and peaks nearly simultaneously with
the AR, but also enhanced its flare-productivity. As a SXR at 12:49:24 UT. It is noted that the first-stage
result, a series of much more energetic flares including rise of the EVE 133 Å/GOES SXR emission is tempo-
two X-class and five M-class ones were produced in the rally correlated to the impulsive rise of the EVE 304 Å
AR during the second period, and furthermore, some of emission. After the EVE 133 Å peak, the cooler coro-
them exhibited an evident EUV late phase. The pro- nal emissions peak sequentially at 12:51:34 UT in EVE
duction of EUV late phase in the X2.1 eruptive flare on 94 Å (Fe XVIII λ93.9 Å, log(T /K) ∼ 6.8, Figure 1(d))
September 6 has been investigated by Dai et al. (2013). and at 12:54:44 UT in EVE 335 Å (Fe XVI λ335.4 Å,
In this work, we focus on an M1.2 non-eruptive flare on log(T /K) ∼ 6.5, Figure 1(e)). The small time lags be-
September 9 when the AR had been decaying. tween these emission peaks indicate a fast cooling pro-
4 Dai et al.

cess in the corresponding flare plasma, whose inferred co-temporal small bumps seen in the GOES and EVE
cooling rate (∼ 2 × 104 K s−1 ) is comparable to those time profiles in Figure 1. The brightening first appears
found in Chamberlin et al. (2012). in a set of sheared arcades enveloping the EW filament
In addition to the first peaks, there exist second con- (highlighted by the arrow in Figure 2(c1)), and then
spicuous emission peaks in both EVE 94 Å and EVE shows a quick eastward propagation along the EW fila-
335 Å, which occur at 13:05:54 and 13:37:34 UT, respec- ment. When the pre-flare brightening propagates to the
tively. The time lag between the second and first peaks NS filament, this filament is destabilized and starts to
in EVE 335 Å is 43 minutes, lying close to the lower ascend rapidly. As the filament rises up, its body ab-
limit of the time delay of a warm coronal late-phase peak sorbs more background coronal emission along the line-
with respect to the corresponding main flare peak (41 of-sight (LOS). As a result, the ascending filament be-
to 204 minutes) in the sample of Woods et al. (2011). comes much more clearly seen as a curved dark feature
The emission ratio between the two peaks is 1.54 ± 0.20 in AIA 335 Å (indicated by the yellow arrow in Fig-
(considering the fluctuations in the original data), nearly ure 2(c2)). Meanwhile, multiple flare ribbons brighten
twice the average value (0.8) in Woods et al. (2011). up quickly in all AIA passbands, coinciding with the
Based on this high peak emission ratio, we adopt the flare’s impulsive phase (the second row of Figure 2). As
same definition by Liu et al. (2015) and hence name this pointed out by the white arrows in Figure 2(c2), the flare
second peak an extremely large EUV late-phase peak. ribbons not only develop from two footpoints of the as-
It is interesting that a large late phase is also evident cending filament, but also appear in some other remote
in EVE 171 Å, whose peak occurs further later (after locations to the east and north of the filament. Near
14:00 UT). the peak of the impulsive phase (∼12:42 UT), several
long brightening loops anchored on the eastern remote
2.2. Correspondence in Spatially Resolved flare ribbon are already observable in the hot passbands
Observations such as AIA 131 (Figure 2(a2)) and 94 Å (Figure 2(b2)),
To resolve activities responsible for the flare emissions, whereas they are invisible in the cooler AIA 335 (Fig-
we inspect AIA images of the flare-hosting AR. AIA ure 2(c2)) and 171 Å (Figure 2(d2)) passbands.
presents multiple simultaneous full-disk images of the The fast rising filament then slows down and finally
TR and corona in 10 passbands with a pixel size of 0.′′ 6 stops at a certain altitude. Material of the filament falls
and a temporal resolution of 12 s (24 s) for the EUV down towards its two footpoints along the spine, making
(UV) passbands. Here we choose a sub-region that cov- the filament eventually drained off (evident in Anima-
ers a field of view (FOV) of 200′′ ×200′′ (x ∈ [610′′ , 810′′ ], tion 1). Coronagraph images from the Large Angle and
y ∈ [80′′ , 280′′]) enclosing the entire AR. The online An- Spectrometric Coronagraph (LASCO; Brueckner et al.
imation 1 presents the flare evolution in six coronal pass- 1995) on board the Solar and Heliospheric Observatory
bands of AIA. Since the evolution behavior of the flare (SOHO ) spacecraft have not recorded any coronal mass
in the AIA 211 and 193 Å passbands is generally simi- ejections (CMEs) during this period, further confirming
lar to that in AIA 171 Å, we only extract characteristic that the flare is non-eruptive. Around the peak time
snapshots in AIA 131 (Fe XXI, Fe VIII), 94 (Fe XVIII, of GOES SXR emission (the third row of Figure 2), in-
Fe X), 335 (Fe XVI), and 171 Å (Fe IX) from the anima- tensely brightening loops appear underneath the previ-
tion, and display them in Figure 2. Note that the dom- ously ascending filament (outlined by the arrow in Fig-
inant ion of each AIA passband is either the same as or ure 2(c3)). In the framework of the CSHKP model, these
of a similar formation temperature to that of the corre- loops correspond to the post-flare loops in the wake of
sponding EVE line selected in Figure 1 (O’Dwyer et al. an erupting flux rope, which exhibit the main phase of
2010). As indicated by the arrows in Figure 2(d1), there the flare. Note that in this row, each frame is taken at
are two low-lying filaments located in the pre-flare AR: the time when the corresponding sub-region intensity at-
one north-south oriented (hereafter NS), and the other tains a local maximum. Hence the time lags between the
east-west oriented (hereafter EW). The filaments are ini- frames in the same row should imply a cooling process
tially most discernible in the cool passbands such as AIA in the main flaring loops. In addition, the long brighten-
171 Å (and also AIA 131 Å because of its low temper- ing loops anchored on the remote flare ribbon, which can
ature sensitivity around 0.6 MK in addition to the high already be seen in the impulsive phase, are now fully de-
temperature response peak over 10 MK). It seems that veloped as revealed in AIA 131 Å (Figure 2(a3)). Com-
the southern end of the NS filament is connected to the pared to the main flaring loops, these longer brightening
eastern end of the EW filament, apparently constituting loops are less bright but cover a much larger spatial ex-
a single L-shaped filament. Nevertheless, the following tent. Nevertheless, they are still invisible in AIA 335
flare evolution shows that these two filaments are indeed (Figure 2(c3)) and 171 Å (Figure 2(d3)).
evolving separately. The main flaring loops quickly fade out, while the long
The precursor of the flare is seen as a pre-flare bright- brightening loops continues to brighten up and reaches
ening in all AIA passbands around 12:30 UT (the first an emission peak sequentially in cooler AIA passbands,
row of Figure 2), which should be responsible for the as shown in the bottom row of Figure 2. Obviously,
EUV LATE PHASE IN A NON-ERUPTIVE FLARE 5

0.5 1.0 1.5 2.0 2.5 3.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.5 1.0 1.5 2.0 2.5 3.0 2.0 2.5 3.0 3.5
log(IAIA/DN s−1) log(IAIA/DN s−1) log(IAIA/DN s−1) log(IAIA/DN s−1)
(a1) AIA 131 Å (b1) AIA 94 Å (c1) AIA 335 Å (d1) AIA 171 Å

Pre−Flare Brightening
250
Pre−Flare
Brightening
Y (arcsec)
Filaments
200

150

100 12:29:57 UT 12:29:50 UT 12:29:51 UT 12:29:50 UT

(a2) (b2) (c2) Flare Ribbons (d2)


250

Impulsive Phase
Ascending
Y (arcsec)

Filament
200

150

100 12:42:21 UT 12:42:14 UT 12:42:15 UT 12:42:14 UT

(a3) (b3) (c3) (d3)


250
Main Flaring Loops

Main Phase
Y (arcsec)

200

150

100 12:50:21 UT 12:51:38 UT 12:54:03 UT 12:57:26 UT

(a4) (b4) (c4) Late−Phase Loops (d4)


250

Late Phase
Y (arcsec)

200

150

100 14:01:21 UT 13:05:02 UT 13:37:39 UT 14:01:12 UT

650 700 750 800 650 700 750 800 650 700 750 800 650 700 750 800
X (arcsec) X (arcsec) X (arcsec) X (arcsec)

Figure 2. Snapshots of the flare evolution in four AIA coronal passbands. The frames are extracted from the online animation
to characterize different phases of the flare evolution, and the arrows highlight some characteristic structures, which are decribed
in detail in the text.
the long brightening loops in AIA 335 Å (indicated by low temperature (∼0.6 MK) to which the AIA 131 Å
the arrows in Figure 2(c4)) are responsible for the ex- passband is also sensitive.
tremely large EUV late phase revealed in EVE 335 Å. We further calculate the sub-region intensities by sum-
Hence we are convinced to identify them as late-phase ming the count rates over all AIA pixels in the FOV
loops. Furthermore, the late-phase loops in AIA 335 Å of Figure 2. The intensity profiles (after subtracting a
(∼13:37 UT, Figure 2(c4)) are morphologically similar pre-flare background averaged from 11:30 to 11:35 UT,
to those seen earlier in the hotter passbands like AIA the same as that chosen in generating the EVE profiles)
131 (∼12:50 UT, Figure 2(a3)) and 94 Å (∼13:05 UT, in five AIA passbands (here we include AIA 304 Å in
Figure 2(b4)), and later in the cooler passband like AIA addition to the four passbands displayed in Figure 2)
171 Å (∼14:01 UT Figure 2(d4)), indicating that they are accordingly overplotted in Figures 1(b)–(f). In each
are the same structures seen in different passbands and panel, the evolution of the AIA sub-region profile is sim-
also at different times. Here the larger time lags be- ilar to that of the EVE profile, with the AIA peaks (out-
tween the appearance times of loops in different AIA lined by the vertical dashed lines) occurring nearly si-
passbands reflect a much slower cooling process in the multaneously with the corresponding EVE peaks (typ-
long late-phase loops than that in the main flaring loops. ically within 1 minute). In particular for the pairs of
As mentioned above, the re-appearance of the late-phase EVE 133 Å/AIA 131 Å, EVE 94Å/AIA 94 Å, and EVE
loops in AIA 131 Å around 14:01 UT (Figure 2(a4)) 335 Å/AIA 335 Å, the Pearson correlation coefficient
means that the late-phase loops have cooled down to a between the two paired profiles is as high as 0.98–0.99.
Moreover, the emission ratio of the late-phase peak to
6 Dai et al.

the main flare peak in AIA 335 Å is 1.36, which is, if 12:47:53 UT (highlighted by the vertical dashed lines),
considering the commonly broader temperature cover- respectively, indicative of two main episodes of energy
age of an AIA passband compared to that of the corre- release taking place in quick succession. With the aid
sponding EVE line, consistent with the ratio revealed in of the profile, we select the AIA 1600 Å snapshots close
EVE 335 Å. We have also changed the size of the sub- to or at the times of the intensity peaks, as displayed in
region used for intensity calculation. It is found that Figures 3(b)–(d). The first-stage energy release is very
as long as the size of the sub-region is large enough to energetic; flare ribbons brighten up promptly, coincid-
contain the whole AR, its variation only marginally af- ing with the flare’s impulsive phase. Around the peak of
fects the resulting AIA profiles. All these factors suggest the ribbon brightening (Figures 3(b) and (c)), the flare
that during this period, the activities taking place inside ribbons are manifested as an elongated one (R1) located
the AR contribute predominantly to the variabilities in in the eastern positive polarity P1 and a semi-circular
full-disk integrated solar emissions in EUV wavelengths, one (R2) located in the western negative polarity N1.
especially for those formed in hot and warm coronal plas- R1 corresponds to the remote flare ribbon seen in Fig-
mas. Since the Fe IX emission mainly comes from the ure 2, whereas R2 is not continuous but broken at its
cool bulk corona, emission fluctuations outside the AR middle into two segments: R2n in the north and R2s in
may significantly influence the overall irradiance in EVE the south. A third flare ribbon is believed to exist in
171 Å, hence making the original data for this line rather the parasitic positive polarity P2. However, saturation
noisy, as opposed to those in other EVE lines. Finally, of the CCD pixels in this region largely masks the rather
it is worth noting that there are a small spike around compact ribbon, making it undistinguishable from the
13:15 UT and a moderate bump after 13:40 UT in the nearby ribbon segment R2s. To validate our specula-
GOES and EVE profiles, which nevertheless have no tion, we reconstruct the Reuven Ramaty High Energy
counterparts in the AIA sub-region profiles. By care- Solar Spectroscopic Imager (RHESSI ; Lin et al. 2002)
fully checking the AIA full-disk images, it is found that image at the energy band of 12–25 keV in the interval of
they are caused by two small brightening events occur- 12:41:54–12:42:18 UT, which reveals two HXR sources
ring in another AR NOAA 11289. (outlined by the red contours in Figure 3(b)) located in
the “masked” compact ribbon and the southern ribbon
2.3. Two-Stage Energy Release segment R2s, respectively. When overplotting the HXR
The evolution of the flare is also traced in the chro- sources onto the simultaneous AIA 335 Å image inserted
mospheric AIA 1600 Å passband, as demonstrated in in Figure 3(b), it is further found that they are excel-
the online Animation 2. Emission in this passband in- lently co-spatial with the two footpoints of the ascending
cludes the continuum formed around the temperature NS filament.
minimum region and the C IV λ1548.2/1550.8 Å dou- Compared to the first-stage energy release, the second-
blet lines formed in the upper chromosphere and TR. stage energy release is more gentle; the flare ribbons
During a solar flare, the optically thin C IV lines are re-brighten up moderately after the first peak. For the
significantly enhanced, showing a prompt and sensitive eastern remote ribbon R1, the brightening is observed
response to the flare energy release. Therefore, the evo- to propagate from north to south along the ribbon. For
lution in AIA 1600 Å can be regarded as a proxy for the western semi-circular ribbon R2, a newly brighten-
flare heating (Li et al. 2012; Qiu et al. 2012; Zhu et al. ing ribbon segment R2m appears and becomes the most
2018). Some snapshots of the animation, as well as an prominent section along this ribbon, filling the gap be-
LOS magnetogram of the AR taken near the time of tween the previously disconnected ribbon segments R2n
flare by the Helioseismic and Magnetic Imager (HMI; and R2s. Meanwhile, the previously masked compact
Scherrer et al. 2012) on board SDO, are displayed in ribbon in the parasitic polarity P2 now becomes clearly
Figures 3(a)–(d). Three main magnetic polarities are discernible, which is identified as R3 (Figure 3(d)). As
identified and labeled as P1, N1, and P2, of which P2 is the energy release goes on, both R2m and R3 experience
a parasitic positive polarity embedded in the host neg- a separation motion away from the polarity inversion
ative polarity N1 (Figure 3(a)). During the course of line (PIL) between them, which is usually observed in a
the flare, flare ribbons are observed in all of these po- two-ribbon solar flare. All these evolution patterns are
larities (Figures 3(b)–(d)), like those seen in Figure 2 clearly revealed in Animation 2.
but free of any possible contaminations from the loop In addition to the time profile of the full AR, in Fig-
structures. The appearance of multiple flare ribbons is ure 3(e) we also plot the time profiles of the AIA 1600 Å
believed to be a natural consequence of the multipolar intensity for regions covering the eastern and western
magnetic field. When overplotting the contours of LOS ribbons, which are outlined in Figure 3(c) by the orange
magnetic field onto the AIA 1600 Å image (Figure 3(c)), and blue boxes, respectively. During the first-stage en-
it is seen that the flare ribbons generally occur in loca- ergy release, the intensities in both regions show a syn-
tions of magnetic concentration. chronous enhancement, with the first peaks in all time
The AIA 1600 Å intensity profile of the AR (purple) profiles occurring at exactly the same time. Neverthe-
shown in Figure 3(e) reveals two peaks at 12:42:41 and less, the second intensity peak for the region of eastern
EUV LATE PHASE IN A NON-ERUPTIVE FLARE 7

−1000 −500 0 500 1000 1.5 2.0 2.5 3.0 3.5


BLOS (G) log(IAIA/DN s−1)
AIA 335 Å 14
(a) HMI LOS (b) AIA 1600 Å 12:42:03 UT (e) 1600 Å Full AR
12 Western Ribbons

AIA Intensity (106 DN s−1)


250 S0
Eastern Ribbon
10
R2n
Y (arcsec)

200 P2 8
P1 R1
6
150 N1 4
R2s
RHESSI 12−25 keV
12:41:54−12:42:18 UT 2
100 12:20:07 UT 12:42:17 UT 0

GOES Flux Deriv. (10−8 W m−2 s−1)


10 3.0

RHESSI Count Rate (103 c s−1)


(c) AIA 1600 Å (d) AIA 1600 Å (f) GOES 1−8 Å
250 8 RHESSI 12−25 keV
2.5
6
2.0
Y (arcsec)

200 S2
R3 4
S1 1.5
S3
2
150 R2m
1.0
0
0.5
100 12:42:41 UT −2
12:47:53 UT
0.0
650 700 750 800 650 700 750 800 12:30 12:40 12:50 13:00 13:10
X (arcsec) X (arcsec) Time (UT on 2011−Sep−09)

Figure 3. Left and middle: HMI line-of-sight (LOS) magnetogram of AR 11283 taken at 12:20 UT (a), and snapshots showing
the flare evolution in AIA 1600 Å (b)–(d), which are extracted from the online animation. The LOS magnetogram reveals
three main magnetic polarities, whose contours are overpoltted in panel (c), with levels of ±200, ±600, and ±1000 G (pink for
positive, and cyan for negative). Multiple flare ribbons (prefixed with the letter “R”) are identified in the AIA 1600 Å images.
The inset in panel (b) is a simultaneous AIA 335 Å image, whose FOV is outlined by the dashed box. Also overplotted in this
panel is a simultaneous RHESSI image at energies of 12–25 keV reconstructed with the Pixon algorithm, with contour levels
corresponding to 30%, 50%, 70%, and 90% of the maximum intensity, respectively. Four slices (S0–S3) are drawn as the yellow
dashed lines or curves in panels (b) and (d), for which the time-distance plots are presented in Figure 4. Right: time profile of
the AIA 1600 Å intensity for different regions (e) and time profiles of soft (derivative) and hard X-rays (f). The vertical dashed
lines indicate the peak times for the corresponding profiles. The AIA intensity is integrated over sub-regions enclosed by the
colored boxes in panel (c).
ribbon is much less discernible than that for the region As its ascending speed gradually approaches zero, the
of western ribbons. It implies that during the second- filament disappears abruptly in the S0 plot, which is
stage energy release, the majority of the released energy due to the falling of material from the filament apex.
should be deposited into the loops anchored on the west- Shortly after the “disappearance” of the filament
ern ribbons. (within 1 minute), the second-stage energy release comes
We use time–distance plots to further study the en- into play. Brightening occurs successively from north
ergy release process. Four slices are selected and drawn to south along the northern section of R1, revealing
in Figures 3(b) and (d), among which S0 extends across an apparent motion speed of 180 km s−1 (Figure 4(b)).
the apex of the ascending filament while S1–S3 lie either Elongation motions of flare ribbon brightening along the
parallelly or perpendicularly to the flare ribbons. The PIL have been both theoretically modeled and obser-
temporal evolution along these slices in AIA 335 Å (S0) vationally seen in solar flares (e.g., Priest & Longcope
and 1600 Å (S1–S3) is shown in Figure 4. A close cor- 2017; Qiu et al. 2017). In this case, the observed slip-
relation between the rise of the filament and the bright- ping of the ribbon brightening is more likely linked to
ening of the flare ribbons is found. A linear fit to the 3D slipping/slip-running reconnection (Aulanier et al.
trajectory of the ascending filament apex reveals a rise 2006, 2007) taking place in a quasi-separatrix layer
speed of 175 km s−1 projected onto the plane of the (QSL; Démoulin et al. 1996). In addition to the appar-
sky (Figure 4(a)). Coinciding with the fast rise of the ent elongation motion of R1, both the newly brightening
filament, the flare ribbons (R1, R2n, R2s, and possi- ribbon segment R2m (Figure 4(c)) and the previously
bly masked R3) brighten up simultaneously and quickly masked ribbon R3 show a separation motion from the
reach the first peaks exactly at the time when the fast as- PIL (Figure 4(d)). Compared to R3, the motion of R2m
cending filament turns to decelerate (Figures 4(b)–(d)). is more reliable to trace. A linear fit to the outer bound-
8 Dai et al.
60
(a) S0 AIA 335 Å NW light curve is that the satellite has orbited into the Earth
Distance (arcsec) 50
shade at this time.

km s −1
40
Ascending
30 Filament

175
20
3. MAGNETIC TOPOLOGY OF THE AR
10
0 SE
Magnetic topology is crucial for understanding the
12:30 12:35 12:40 12:45 12:50 12:55 13:00
Peak 1 Peak 2 magnetic reconnection process in a solar flare and the
100 (b) S1 AIA 1600 Å N consequent morphological evolution of the flare loops
and ribbons. In a typically low β (the ratio of gas pres-

18
80

0k
sure to magnetic pressure) corona, we explore the mag-

m
60

s
R1 netic topology of the flare-hosting AR with the aid of

−1
40

20
an NLFFF extrapolation.
0 S We adopt the optimization method proposed by
120 (c) S2 AIA 1600 Å N Wheatland et al. (2000) and implemented by Wiegelmann
(2004) to perform the NLFFF extrapolation. To pre-
Distance (arcsec)

100
R2n R2m
80 pare the bottom boundary, we use an HMI vector mag-
60
R2s netogram of the AR taken close to the time of flare
40
occurrence (∼11:58 UT), which is selected from the
20
0 S Space-weather HMI Active Region Patches (SHARPs).
30
(d) S3 AIA 1600 Å W The pipeline to generate SHARP data products includes
25 R3
the removal of the 180◦ ambiguity in transverse field
20
(Metcalf et al. 2006; Leka et al. 2009), and the remap-
15
20 ping of the magnetic field vector (Gary & Hagyard 1990)
10 km
5
s −1 in a cylindrical equal-area (CEA; Calabretta & Greisen
0 R2m E 2002) projection (for more details, see Bobra et al. 2014
12:30 12:35 12:40 12:45 12:50 12:55 13:00 and Hoeksema et al. 2014). The grid size used for the
Time (UT on 2011−Sep−09)
SHARP sampling is 0.03◦ in heliocentric angle, corre-
Figure 4. Time-distance plots along the selected slices in sponding to a pixel size of ∼0.36 Mm. To save the
AIA 335 Å (a) and 1600 Å (b)–(d), within which some char- computation time while not significantly losing the es-
acteristic structures are identified and labeled. Linear fit- sential features in the magnetic field modeling, we ex-
tings are applied to some moving structures traced by the tract a sub-region that is large enough to cover all main
yellow dotted lines, with the fitted speeds labeled along these polarities in the original SHARP magnetogram, and
lines. Overplotted in panels (a) and (d) is the AIA 1600 Å
degrade the sampling resolution by a factor of 2. A
further preprocessing is applied to the area of interest
intensity profile of the AR (the same as that plotted in Fig-
to minimize the net magnetic force and torque on the
ure 3(e)) revealing two peaks by the vertical dashed lines in
boundary (Wiegelmann et al. 2006). The extrapolation
all panels.
is then conducted in a box of 300 × 200 × 200 grid points
uniformly spaced in the computation domain.
ary of R2m reveals a separation speed of 20 km s−1 . To The magnetic topology of the AR based on our
compensate for the projection shortening at the flare NLFFF extrapolation is presented in Figure 5. The
position of ∼N13◦ W52◦ , we multiply the fitted speed most prominent features are a 3D magnetic null point
by a factor of ∼1.6, and obtain a corrected separation located at a height of 11.8 Mm, and inner (green)
speed of ∼30 km s−1 , which is consistent with the results and outer (blue) fan-spine magnetic field lines pass-
previously found in two-ribbon flares (e.g., Asai et al. ing around it (Lau & Finn 1990). With the aid of the
2004; Miklenic et al. 2007). normal component (Bz ) of the SHARP vector magne-
The two episodes of energy release are also evident togram (Figures 5(a)–(c)), it is seen that the footpoints
in X-rays. Figure 3(f) shows the derivative of the of the inner and outer spine field lines reside in the
GOES 1–8 Å flux (black), which is generally taken as parasitic and eastern positive polarities P2 and P1,
a proxy for the HXR emission owing to the “Neupert respectively, while the fan field lines form a dome-like
effect” (Neupert 1968). Two peaks are readily resolved structure, whose base intersects with the solar surface in
at 12:42:33 and 12:47:45 UT (indicated by the vertical the negative polarity N1. Embedded in the fan dome,
dashed lines), respectively, in perfect coincidence with two magnetic flux ropes (red for the NS-oriented one
the corresponding AIA 1600 Å peaks (within 10 s). The and yellow for the EW-oriented one) are found to ex-
overplotted RHESSI 12–25 keV light curve (red) reveals ist. Through co-alignment with the AIA coronal images
only one peak at 12:42:26 UT, which coincides with the (not shown here), it is found that the flux ropes are ap-
first peak in the GOES derivative curve. The reason proximately co-spatial with the two low-lying filaments
for the absence of a second peak in the RHESSI HXR identified before the flare.
EUV LATE PHASE IN A NON-ERUPTIVE FLARE 9

4 (a) NLFFF (Top View) 4 (d) NLFFF (Top View)


2000

2 1000 2
0 (G)
Y (deg) 0 0

Y (deg)
−1000
−2 P1 −2000 −2 R1
P2 R3
N1 R2
−4 −4

−6 −6
Bottom: SHARP BZ 11:58:22 UT Bottom: AIA 1600 Å 12:47:53 UT (CEA)
−5 0 5 10 −5 0 5 10
X (deg) X (deg)
1 (b) NLFFF (Top View) 1 (e) NLFFF (Top View)
5
0 0 4

−1 −1 3
Y (deg)

Y (deg)
2
−2 −2
1
−3 −3
Outer Fan−Spine
−4 Inner Fan−Spine −4
NS Flux Rope
−5 EW Flux Rope −5 Bottom: logQ (z=0)
0 2 4 6 8 0 2 4 6 8
X (deg) X (deg)
1 2 3 4 5
(c) NLFFF (Side View)
(f) logQ (Vertical)
25
Spine
20
Z (Mm)

15
10 Fan
5
0
0 10 20 30 40 50 60
Distance (Mm)

Figure 5. NLFFF extrapolated magnetic field lines showing fan-spine and flux rope structures, which are overlaid on the
boundary SHARP Bz magnetogram (a)–(c), an AIA 1600 Å CEA map (d), and a filtered map of the squashing factor Q on
the bottom boundary (e), respectively. Also shown is a vertical Q-map revealing a fan-spine QSL structure (f). The field lines
are color-coded according to the legends in panel (b), and the Q-map is shown for the vertical cut along the orange dashed line
plotted in panel (e). In the left column, panels (b) and (c) show zoomed-in top and side views of the field lines, respectively.
Note that the coordinate origin in the top-view panels corresponds to the projection center of the original SHARP CEA map,
which has a Carrington coordinate of (220.28◦ , 17.39◦ ).
In Figure 5(d) we overlay the above magnetic field the chromospheric flare ribbons. The footpoints of the
lines on an AIA 1600 Å image taken at the second peak inner and outer spines are located exactly in the com-
in this passband (12:47:53 UT) when all the flare ribbons pact ribbon R3 and the eastern remote ribbon R1, re-
are clearly seen. For a better comparison between them, spectively. The degree of match found here seems bet-
we remap the AIA 1600 Å image into the same CEA ter than those in previously reported cases where the
projection as that of the SHARP map. Note that such spatial discrepancy between the outer spine footpoint
remapping will introduce artificial distortions to above- and the remote flare ribbon is typically over 10 Mm
the-surface structures, especially for those located far (e.g., Vemareddy & Wiegelmann 2014; Yang et al. 2015;
away from the disk center, just like the late-phase loops Masson et al. 2017). Except for the northernmost sec-
observed in this flare. For this reason, we do not at- tion, the circular ribbon R2 also matches the base of the
tempt to compare the extrapolated magnetic field with fan dome very well. In addition, the two footpoints of
the flare loops seen in other AIA coronal passbands. As the NS flux rope are anchored at ribbons R2 and R3,
revealed in Figure 5(d), there is a good match between respectively.
the footpoints of the overlaid magnetic field lines and
10 Dai et al.

The null point and the fan-spine structure indicate previous case studies, estimation of the lengths of late-
the presence of magnetic separatrix and QSLs, whose phase loops was mainly based on single-perspective ob-
distribution can be quantified by the so-called squash- servations and therefore subject to large uncertainties.
ing factor Q (Titov et al. 2002). The Q factor quantita- Through imaging observations from multiple perspec-
tively characterizes the connectivity gradients between tives, we could in principle track the realistic coronal
one magnetic field line and its neighboring field lines. loops in 3D, and hence more accurately determine the
Here we use the method proposed by Pariat & Démoulin loop lengths.
(2012) to compute the distribution of Q value in the 3D Figure 6(a) depicts the deployment of the Solar Ter-
domain. In Figure 5(e) we plot the distribution of log Q restrial Relations Observatory (STEREO ; Kaiser et al.
on the bottom boundary (z = 0). For the sake of clar- 2008) twin spacecraft with respect to the Earth at the
ity, we apply a filter to the Q-map so that only the time of flare. The longitudinal position of the flare site
distribution of log Q in regions of |Bz | >100 G is dis- (indicated by the black dot) and its favorable viewing
played. When overlaying the modeled field lines on the angles to STEREO-A and near-Earth SDO enable us
Q-map, it is clearly seen that the fan-spine field lines to stereoscopically measure the 3D coordinates of flare
intersect with the solar surface at patches of high log Q loops captured by both satellites. In Figures 6(b) and
values. This is consistent with the observational fact (c) we display paired images of the late-phase loops
found in many previous studies (e.g., Chen et al. 2012; taken simultaneously at ∼13:50 UT by the EUV Im-
Masson et al. 2017) that flare ribbons typically stop at ager (EUVI; Wuelser et al. 2004) on board STEREO-A
the intersection of QSLs with the solar surface. To de- (hereafter EUVI-A) in 195 Å and by AIA in 193 Å,
pict the structure of the QSLs in the corona, we draw a respectively, both of which have a similar temperature
slice across the fan-dome base and the footpoint of the response peaking around 1.3 MK. As shown in the fig-
inner spine on the bottom boundary (the orange dashed ures, the late-phase loops appear mainly in two clusters.
line in Figure 5(e)), and plot the 2D Q-map in the ver- In the north, the late-phase loops are readily resolved
tical plane extending above the slice, which is shown in in both passbands. Using the Solar Software (SSW;
Figure 5(f). As expected, it clearly shows the fan-spine Freeland & Handy 1998) routine scc measure.pro, we
magnetic topology of this event, i.e., a QSL surrounding trace a full late-phase loop, whose 3D coordinates are
the fan-spine separatrix (Masson et al. 2009). projected onto the two perspectives (dashed lines), re-
There are still some discrepancies in morphology be- spectively. By integrating the distance along the loop,
tween the extrapolated magnetic field and the flare rib- we derive a length of 81 Mm for this late-phase loop.
bons. First, the northernmost section of the circular rib- In the south, the triangulation measurement of the late-
bon R2 extends well beyond the fan-dome base, showing phase loops is nevertheless significantly affected by an
a spatial offset up to 15 Mm (Figure 5(d)). Second, in LOS effect. When viewed from the two different per-
spite of the excellent match between the outer spine foot- spectives, either the eastern (from EUVI-A) or the west-
point and the eastern remote ribbon R1, field lines that ern (from AIA) legs of the late-phase loops coincidently
originate from other parts of this elongated ribbon, pass overlap along the corresponding LOS. As a result, only
through the fan-spine QSL, and end up at the circular the legs on the opposite side are distinguishable from the
ribbon R2 cannot be reasonably figured out, although corresponding perspective, as pointed out by the arrows
their existence can be confirmed by the appearance of in the figures. Moreover, the intensity of the loop-top
the widespread late-phase loops. For this reason, we do regions in both passbands is predominately high, since
not display these field lines in Figure 5. As mentioned the intensity observed there is actually an integration
above, at the time of flare, the AR is located far away along the LOS, which is also contributed by emissions
from the disk center. The SHARP remapping procedure from the loop legs. Restricted by the LOS effect, here
could hence amplify the errors of magnetic field mea- we just stereoscopically estimate the apex location of
sured in the local plane, which are then transferred to the late-phase loops (indicated by the asterisks), whose
all the field components at the bottom boundary used height is ∼35 Mm above the solar surface. Assuming a
for extrapolation. This may significantly compromise semi-circular loop geometry, we obtain a loop length of
the fidelity of magnetic field extrapolation from regions ∼110 Mm. Note that due to the uncertainty in deter-
of relatively weak field strengths. mining the loop apex, this value might be an underesti-
mation of the lengths for the southern late-phase loops.
4. MODELING OF THE LATE-PHASE LOOPS
The projections of the axis of the late-phase loop in
4.1. Stereoscopic Measurement of the Loop Lengths the north and apex of the late-phase loop in the south
By combing the full-disk integrated EVE and spatially are further overplotted in an AIA 335 Å difference image
resolved AIA observations, a clear correspondence be- (Figure 6(d)) taken 13 minutes earlier (at the late-phase
tween the late-phase emissions and the long late-phase peak in AIA 335 Å). We find an excellent match between
loops in this flare has been validated. The length of the loop geometries and emission features in the two AIA
a coronal loop is, both theoretically and observation- passbands, further confirming the slow cooling process
ally, an important parameter for the loop evolution. In in the long late-phase loops.
EUV LATE PHASE IN A NON-ERUPTIVE FLARE 11

(a) (b) EUVI−A 195 Å


−1.0

250
−0.5
A
HEE_Y (AU)

Y (arcsec)
B Sun 200
0.0
Flare

0.5 150
Western Legs

1.0
Earth 100 13:50:30 UT

−1.0 −0.5 0.0 0.5 1.0 −850 −800 −750 −700


HEE_X (AU) X (arcsec)

(c) AIA 193 Å (d) AIA 335 Å

250 250

200 200
Y (arcsec)

Y (arcsec)

150 150
1 2 1 2

Eastern Legs Eastern Legs


100 100
13:50:43 UT 13:37:39 UT

650 700 750 800 650 700 750 800


X (arcsec) X (arcsec)

Figure 6. Deployment of the STEREO twin spacecraft with respect to the Earth at the time of flare (a), and images of the
late-phase loops viewed from two different perspectives (b)–(d). In panel (a), the black dot indicates the longitudinal position of
the flare site. In panels (b)–(d), the dashed lines and asterisks outline the projection of the characteristic late-phase loops onto
the two different perspectives, respectively, and the arrows point to the legs of the southern late-phase loops. In the bottom
row images, two boxes are drawn at the leg and footpoint of a southern late-phase loop, respectively, over which we derive the
AIA light curves used for modeling of the late-phase emissions.
For comparison, we also quantify the lengths of the 4.2.1. Model Setup
main flaring loops, which are simply inferred from the
separation distance between the flare ribbons R2m and The EBTEL model is a zero-dimensional (0D) hydro-
R3 on which the two conjugate loop footpoints are an- dynamic model that describes the evolution of the aver-
chored. The estimated lengths of the main flaring loops age temperature, pressure, and density along a coronal
are 10–25 Mm. Obviously, the two sets of flare loops loop (or more strictly speaking, a coronal strand). The
are distinct in length, with a length ratio between the basic idea behind EBTEL is that the imbalance between
late-phase loops and the main flaring loops exceeding 3. the heat flux conducting down into the TR and the ra-
diative loss rate there leads to an enthalpy flux at the
coronal base: an excess heat flux drives an evaporative
4.2. EBTEL Modeling of the Late-Phase Emissions upflow, whereas a deficient heat flux is compensated for
The electromagnetic emissions from a flaring loop re- by a condensation downflow. This upward/downward
flect the underlying evolution of temperature and den- enthalpy flow controls the hydrodynamic evolution of
sity, and thereby the energy release history in the loop. the coronal loop.
Using the EBTEL model, we model the EUV emissions Another assumption of EBTEL is that any flows along
of a late-phase loop in this flare. By comparing the syn- the loop are subsonic, which should hold for most of
thetic EUV light curves with the observed light curves the time of the loop evolution. This means that the
in several AIA passbands, we quantify parameters of the kinetic energy is negligible compared to the internal en-
heating on the late-phase loop as well as the loop geom- ergy. Hence the equation of momentum can be dropped
etry. from the governing equations, so do the kinetic terms
12 Dai et al.

Norm. AIA Intensity


1.0 (a) 1600 Å
from the equation of energy conservation. Following this Observed
Gaussian Fit
0.8 Box 1
way, in EBTEL we solve the remaining loop-averaged 0.6
t0=12:43:08 UT
τ=23 s
(along the coronal section of the loop) hydrodynamic 0.4
equations including the equations of continuity and en- 0.2
0.0
ergy conservation, and equation of state, which are given
12:40 12:44 12:48 12:52 12:56 13:00
by 1000
dn c2 (b) 131 Å
800 Box 2
Observed
=− (F0 + c1 Rc ), (1) Synthetic
dt 5c3 kT L 600
  400
dp 2 Rc 200
= H − (1 + c1 ) , (2)
dt 3 L 0
250
(c) 94 Å
and 200 Box 2
1 dT 1 dp 1 dn 150

AIA Intensity (DN s−1 pix−1)


= − , (3)
T dt p dt n dt 100
50
where n, p, and T are the average density, pressure, 0
and temperature of the coronal loop, respectively, L is 120 (d) 335 Å H0=1.1 erg cm−3 s−1
100 Box 2 L=82 Mm
the loop half-length (EBTEL assumes a symmetric semi- 80
WL=5.2 Mm
circular loop geometry and hence treats only half of the 60
40
loop), k is the Boltzmann constant, c2 (c3 ) is the ra- 20
tio of the average coronal (coronal base) temperature to 0
apex temperature, F0 = −(2/7)κ0 (T /c2 )7/2 /L (where 3000 (e) 171 Å
Box 2
κ0 = 8.12×10−7 in cgs units is the classical Spitzer ther- 2000
mal conduction coefficient) is the heat flux at the coronal
1000
base, Rc = n2 Λ(T )L (where Λ(T ) is the optically thin
radiative loss function) approximates the radiative loss 0

rate from the corona, c1 = Rtr /Rc is the ratio of radia- 12:40 13:00 13:20 13:40 14:00 14:20 14:40
Time (UT on 2011−Sep−09)
tive loss rate of the TR to that of the corona, and H
denotes the average volumetric heating rate (note that Figure 7. Top: Normalized time profile observed at the
here we use H instead of Q, the commonly adopted nota- footpoint of the selected late-phase loop (Box 1 in Figure 6)
tion for heating rate in the literature, to avoid confusion in the AIA 1600 Å passband (solid curve), with the Gaus-
with the squashing factor introduced in Section 3). sian fit overplotted (dashed curve). Bottom: Light curves
Of the three dimensionless parameters c1 , c2 , and c3 observed at the leg of the same loop (Box 2 in Figure 6)
in EBTEL, c1 is the most important. The main dif- in four AIA coronal passbands (solid curves), as well as the
ference between the different versions of EBTEL lies in synthetic light curves (dashed curves) computed with the
the treatment of c1 . Assuming a single power-law radia-
best-fit parameters in the EBTEL modeling (specified in the
tive loss function, Martens (2010) analytically studied
legends in panel (d)).
the energy equilibrium for a static coronal loop, whose
solutions reveal a value of c1 being around 2, and val-
ues of c2 and c3 close to 0.9 and 0.6, respectively. In at the same temperature, leading to higher values of c1
the original EBTEL model (Klimchuk et al. 2008), to (Barnes et al. 2016). During the radiative cooling phase,
achieve an overall consistency with the 1D simulation the loop is instead over-dense, which in turn reduces the
results, c1 was kept as a constant of 4 throughout the c1 values (Cargill et al. 2012). In our modeling, we hold
loop evolution. Nevertheless, when a loop evolves dy- c2 and c3 at typical values of 0.9 and 0.6, respectively, as
namically, the equilibrium assumption is broken and c1 specified in all EBTEL versions, whereas we use the lat-
also evolves. Obviously, the choice of a fixed c1 value is est EBTEL version (Barnes et al. 2016) to consistently
not physically reasonable. In the later EBTEL versions, calculate the c1 values.
more physics was included to appropriately determine
4.2.2. Comparison with Observations
the values of c1 . One piece of physics previously ig-
nored is gravitational stratification, whose main effect Based on the output temperature and density from
is to depress the coronal radiation. Thus, higher val- the EBTEL model, we can synthesize the emissions from
ues of c1 can be expected, especially for long loops with the loop. The synthetic intensity of the loop in an AIA
large ratios of the length to the gravitational scale height passband can be computed as
(Cargill et al. 2012). Another piece is the deviation of IAIA = R(T )n2 WL (DN s−1 pix−1 ), (4)
the dynamically cooling loop from equilibrium states.
In the early conductive cooling phase, the density in- where R(T ) is the temperature response function of the
crease in response to the strong heat flux is not so fast AIA passband, which can be obtained by the SSW rou-
that the loop is under-dense with respect to a static loop tine aia get response.pro, and WL is the nominal
EUV LATE PHASE IN A NON-ERUPTIVE FLARE 13

width (depth) of the loop along the LOS. In gener- In our modeling, we only fit the observed light curve of
ating the AIA response functions, we set the evenorm the late-phase loop in AIA 335 Å, since we mainly focus
keyword, which will give better agreement with full-disk on the EUV late phase in the warm coronal emissions.
EVE observations when cross-checking data from both For other passbands, we manually inspect the degree of
AIA and EVE. Due to the 0D nature of the EBTEL match between the synthetic light curves computed with
model, it is physically not applicable to give a spatial the best-fit parameters and the observed light curves.
distribution of the synthetic intensities along the loop. Some factors that could affect the loop emissions in these
To compare the synthetic intensities with observations, passbands are addressed later in this section.
we select a late-phase loop in the south whose observed The fitting gives best-fit values for the three free
intensities are roughly uniformly distributed along its parameters: H0 =1.1 erg cm−3 s−1 , L=82 Mm, and
eastern leg, as shown in the bottom row AIA images WL =5.2 Mm. To test the robustness of the fitting, we
of Figure 6. We draw a 2′′ × 2′′ box over the loop leg vary the starting guesses for the parameters within a
(Box 2 in Figure 6). The intensities averaged over this reasonable range. In most cases, the fitting procedure
box in several AIA coronal passbands are taken as the converges to the same results as those given above. The
observed loop intensities, whose time profiles are shown synthetic light curves computed with the best-fit pa-
as the solid curves in Figures 7(b)–(e). rameters are overplotted as the dashed curves in Fig-
To infer the heating function for the loop, we calcu- ures 7(b)–(e). As seen in Figure 7(d), the peak in both
late the chromospheric light curve at the loop footpoint, the synthetic and the observed light curves in AIA 335 Å
as proposed by Li et al. (2012), Qiu et al. (2012), and occurs almost one hour after the impulsive heating. In
Zhu et al. (2018). To this end, we draw a 6′′ × 6′′ box the other passbands, the peak times in the synthetic
(Box 1 in Figure 6) at the eastern footpoint of the se- light curves also reasonably agree with those in the ob-
lected late-phase loop, which is located on the remote served light curves. This indicates that even without
flare ribbon R1. In Figure 7(a) we plot the normalized an additional delayed heating, the long cooling process
AIA 1600 Å intensity profile averaged over this box (the of the late-phase loop can sufficiently explain the pres-
solid curve), which closely resembles the intensity profile ence of the late-phase emissions at different tempera-
for the whole eastern ribbon shown in Figure 3(e). By tures. Note that there are three consecutive peaks in
applying a Gaussian fit to the rise and fast decay phase the observed AIA 171 Å light curve. It means that the
of the AIA 1600 Å light curve, we derive a peak time at LOS should intersect at least three late-phase loops of
12:43:08 UT and a width of 23 s for the fitted Gaussian similar half-lengths, and the parameter WL in the mod-
function (the dashed curve in Figure 7(a)). Mimicking eling would reflect a combination of the widths of all
the shape of the fitted Gaussian function, the imposed these loops seen along the LOS. In this sense, the ef-
impulsive heating on the late-phase loop is then given fective width of an individual late-phase loop is of the
by order of 1.7 Mm, which is consistent with those found
in previous observations (e.g., Aschwanden et al. 2008).
(t − t0 )2
 
H(t) = H0 exp − (erg cm−3 s−1 ), (5) In spite of a good match between the modeled results
2τ 2 and observations seen in AIA 335 Å, there is a large
where H0 is the amplitude of the heating, t0 =12:43:08 UT discrepancy in amplitude between the synthetic light
the time of the peak heating rate, and τ =23 s the Gaus- curves and the observed ones in the other passbands. In
sian duration of the heating. Note that there is a second the hotter AIA 131 and 94 Å passbands, the synthetic
small peak ∼5 minutes after the first peak in the AIA peak intensities are significantly higher than those ob-
1600 Å intensity profile, which possibly reflects a sec- served. The hotter the passband, the more prominent
ond heating process. Nevertheless, this second heating such a discrepancy (Figures 7(b) and (c)). We attribute
should be much weaker than the first heating, and hence this discrepancy to a possible deviation from an ioniza-
can be safely ignored in our modeling. In addition to the tion equilibrium in the late-phase loops, a factor not
impulsive heating, we also include a steady background considered in our EBTEL modeling. During the im-
heating whose rate is fixed to be 10−5 erg cm−3 s−1 . pulsive heating and conductive cooling stage, the loop
While maintaining the loop temperature and density temperature evolves so fast that the ionization process
before and long after the impulsive heating, this back- cannot catch up with the temperature change, result-
ground heating does not obviously affect the modeled ing in a non-equilibrium ionization (Reale & Orlando
EUV emissions from the loop. 2008; Bradshaw & Klimchuk 2011). Obviously, this ef-
There are three free parameters in our modeling, i.e., fect is the most prominent for hot coronal emissions from
H0 , L, and WL . Although the characteristic loop length long tenuous loops, e.g., the late-phase loops. Using
for the southern late-phase loops has been estimated us- the full 1D HYDRAD code (Bradshaw & Mason 2003),
ing the triangulation method, the LOS effect may intro- Bradshaw & Klimchuk (2011) numerically studied the
duce some uncertainties. Therefore, here we still regard effect of non-equilibrium ionization on loop emissions,
L as a free parameter. We use the MPFIT (Markwardt and found that the synthetic loop intensities in hot pass-
2009) package distributed in SSW to perform the fitting. bands such as AIA 131 Å can be depressed by orders of
14 Dai et al.

magnitude compared to those under the ionization equi- (2012), the modeled loop intensities in AIA 211 and
librium assumption. They also pointed out that as the 171 Å are also systematically lower than the observed in-
cooling rate slows down and the loop density increases, tensities (Figure 4 in their paper). We note that Li et al.
the ions can have enough time to reach an ionization (2012) fixed the parameter c1 for each individual loop,
equilibrium at a medium temperature like 6 MK. There- whose value would be too high for the radiative cooling
fore, emissions formed below this temperature (like the phase.
EUV late-phase emission in AIA 335 Å) are unlikely to
be notably affected by this effect. Another piece of evi- 5. DISCUSSION
dence for the existence of non-equilibrium ionization in Based on the observations and the NLFFF extrapo-
the modeled late-phase loop is the small bump seen in lation results, we propose a two-stage magnetic recon-
the synthetic AIA 335 Å light curve shortly after the im- nection scenario to explain the observed flare evolution.
pulsive heating, which is also synchronous with the peak The flare is preceded by a brightening first occurring
of the synthetic light curve in AIA 131 Å. This bump around or within the EW flux rope (manifested by the
reflects a high temperature response over 10 MK in AIA EW filament), which then propagates eastward and trig-
335 Å, which is introduced by crosstalk from the AIA gers an instability of the nearby NS flux rope (mani-
131 Å passband (Boerner et al. 2012). Nevertheless, due fested by the NS filament). The details of the trigger-
to the possible non-equilibrium ionization at flare tem- ing process are beyond the scope of this work. Here
peratures early in the long tenuous late-phase loop, the we conjecture that the helical kink instability may be
bump is absent in the observed AIA 335 Å light curve. responsible for the fast ascent of the initially low-lying
Recently, Zhu et al. (2018) studied a C-class two-ribbon NS flux rope. As the flux rope rises up, the first-stage
solar flare using the EBTEL model, in which the mod- magnetic reconnection sets in. Inside the flux rope,
eled peaks in AIA 131 and 94 Å are also well above the possible reconnection due to an internal kink instabil-
observed peaks (Figure 3 in their paper). ity (Galsgaard & Nordlund 1997) can heat the filament
On the other hand, the synthetic loop intensities in body, and more importantly, accelerate high-energy par-
the cooler AIA 171 Å passband are systematically lower ticles that produce the two conjugate HXR sources at
than the observed intensities (Figure 7(e)). It implies the footpoints of the flux rope, as proposed in Guo et al.
an over-estimation of the mass draining from the corona (2012). Meanwhile, the flux rope pushes the overlying
during the radiative cooling phase. The reason may lie field lines toward the null point, and makes the fan-
in the geometry of the late-phase loop. We note that the spine QSL structure thinner, causing an enhancement
best-fit value of parameter L for the southern late-phase the current inside it. Reconnections both around the
loop is 82 Mm in the EBTEL modeling, which is signif- null point and in the QSL can produce the observed mul-
icantly higher than the value of ∼55 Mm inferred from tiple flare ribbons and long late-phase loops connecting
the height of the loop apex based on the semi-circular the eastern remote ribbon R1 and circular ribbon R2.
loop assumption. Such a large discrepancy in length is Considering the elongation of the remote ribbon R1 and
unlikely to be simply ascribed to the uncertainty in de- the appearance of the late-phase loops, the QSL recon-
termining the loop height, while it is more likely due to nection seems more energetic than the null-point recon-
an oblate shape of the late-phase loop rather than the nection. Moreover, since all flare ribbons brighten up
semi-circular geometry as commonly assumed. With the very impulsively, we suggest that the QSL reconnection
same loop length, a semi-circular loop can reach a higher could belong to a super-Alfvénic slip-running reconnec-
altitude than an oblate loop. Therefore, in a semi- tion (Aulanier et al. 2006), during which the inferred
circular loop, the coronal emission is more significantly elongation motion of ribbon brightening has been com-
depressed by the gravitational stratification, which ac- plete within the time cadence of AIA.
cordingly leads to a higher c1 value. Since the EBTEL As the fast ascending flux rope slows down, the
model assumes a semi-circular loop geometry for sim- second-stage magnetic reconnection starts to be ini-
plicity, it will overestimate the c1 values when dealing tiated. Reconnection mainly takes place in the current
with an actually oblate-shaped loop, and hence inap- sheet formed between the two legs of the overlying field
propriately elevate the mass draining rate during the lines stretched by the flux rope. The newly formed rib-
radiative cooling phase in order to compensate for the bon segment R2c fills the gap in the circular ribbon R2
over-estimated TR radiative loss (Cargill et al. 2012). formed in the first stage. Short flaring loops connect
As a result, the modeled loop intensities during this R2c with the compact ribbon R3. Such a reconnection
stage will be under-estimated, as is seen in AIA 171 Å is predicted by the standard 2D CSHKP flare model,
for the late-phase loop modeled here. At higher tem- and the observed separation speed of the flare ribbon
peratures, for example, those corresponding to the AIA (∼30 km s−1 after correction for the projection effect)
335 Å passband, the relative mass draining rate (quan- is also typical for a two-ribbon flare. Therefore, we refer
tified as d ln n/d ln T ) is not so fast that the relevant to the observed flaring loops as “main flaring loops”.
loop emission would be much less impacted. In another Meanwhile, there also exists a less energetic QSL recon-
EBTEL modeling of an M-class solar flare by Li et al. nection, which is responsible for the elongation motion
EUV LATE PHASE IN A NON-ERUPTIVE FLARE 15

of small brightening along the remote ribbon R1. The GOES classes (e.g., Li et al. 2012; Zhu et al. 2018), it
revealed motion speed of 180 km s−1 implies that the is found that the peak heating rate of the late-phase
QSL reconnection during this stage may correspond loops in this flare is comparable to or even greater than
to a sub-Alfvénic slipping reconnection (Aulanier et al. the heating rate of the main flaring loops in those flares.
2006). Considering the larger length of the late-phase loop, it is
Based on the two-stage reconnection scenario, the long suggested that the late-phase loop undergoes an intense
late-phase loops are mainly heated during the first-stage heating with a peak energy deposition rate (2H0 L) as
QSL reconnection, while the second-stage flare reconnec- high as 1.8×1010 erg cm−2 s−1 and total energy input
tion, occurring ∼5 minutes later, is responsible for the of 7.1×1028 erg.
heating of short main flaring loops. This process cannot For a late-phase loop with a half-length of 82 Mm,
be accounted for by the standard CSHKP flare model, in the loop emission in AIA 335 Å peaks almost one hour
which shorter flare loops are always heated earlier than after the impulsive heating. Since there is no obvious
longer loops. Nevertheless, it is a natural consequence additional heating on the late-phase loop, the time de-
of a multipolar magnetic topology, as is the case in this lay of the warm coronal emission peak is mainly due
flare. We note that our scenario is basically in agreement to the long cooling process of the late-phase loop, as
with the picture proposed by Sun et al. (2013) based on proposed in Liu et al. (2013). By comparison, the main
a toy model (Figure 1 in their paper) except that in their flaring loops are significantly shorter (2L∼10–25 Mm),
opinion the late-phase emission mainly comes from the and hence evolve more quickly. In a numerical experi-
outer spine lines and is caused by reconnection at the ment using the EBTEL model, Dai & Ding (2018) found
fan-spine null point. that with a strong initial loop heating, the hot and warm
The close correlation between the rise of the flux rope coronal emissions peak in the conductive cooling and ra-
and the brightening of the flare ribbons indicates that diative cooling stages, respectively. For the hot coronal
the flux rope should be the main driving agent for the emissions, the main flare peak and the late-phase peak
reconnections in both stages. During the first stage, the temporally overlap because the cooling rate in all loops
fast ascending flux rope makes the above QSL recon- is very fast during the conductive cooling phase. How-
nection more enhanced than usual. Nevertheless, this ever, the small difference in cooling rate between the
QSL reconnection does not efficiently remove the strong loops of different lengths can still produce an observable
magnetic constraints above the flux rope, which there- bump following the main peak, which results in a dual-
fore slows down and finally fails to erupt. The field lines decay behavior in the corresponding light curves, as seen
involved in the second-stage reconnection have not yet in this flare (Figures 1(a) and (c)). We note that this
been sufficiently stretched when the flux rope stops ris- dual-decay in GOES SXRs has been used as a proxy
ing, causing a significant depression of the reconnection for EUV late-phase flares prior to the SDO era (Woods
rate in this stage. Therefore, it is likely that more en- 2014). When the loops continue to cool down, the effect
ergy is injected to the late-phase loops than the main of the difference in loop length on the loop cooling time
flaring loops, leading to an extremely large EUV late becomes more and more prominent. At warm coronal
phase in this non-eruptive flare. temperatures, the late-phase peak can be well separated
We also quantify the heating history in a typical late- and thus easily distinguished from the main flare peak.
phase loop. The chromospheric AIA 1600 Å light curve Since the warm coronal emission peaks in the radiative
observed at the loop footpoint further confirms that the cooling stage, the emission after the peak can be notably
late-phase loop is heated by an impulsive energy re- affected by the mass draining, hence showing a fast de-
lease during the first-stage reconnection. By using the cay compared to the relatively gradual rise. This pattern
EBTEL model, we synthesize the loop intensities and is consistent with the shape of the light curves of AIA
compare them with observations. Fitting the observed 335 Å in this flare, both for the whole AR (Figure 1(e))
intensities with the synthetic ones yields best-fit values and for the individual late-phase loop (Figure 7(d)).
for the peak heating rate H0 =1.1 erg cm−3 s−1 and In the EBTEL modeling of the late-phase loop, the
loop half-length L =82 Mm. Due to a contamination synthetic peak intensities in the hotter AIA passbands
from the TR and chromosphere, we cannot reliably ex- are significantly higher than those observed. We at-
tract the loop intensities for the much shorter main flar- tribute this discrepancy to a non-equilibrium ionization
ing loops. Furthermore, we cannot reliably characterize effect (Reale & Orlando 2008; Bradshaw & Klimchuk
the contribution of nonthermal electron beam heating, 2011) in the late-phase loop. To further verify this hy-
which is believed to be marginal in the heating of late- pothesis, we perform a new set of fitting in which the
phase loops but becomes important during main flare observational data from both AIA 335 and 94 Å are
heating. These factors prevent us from applying the fitted, as done in Li et al. (2012). In this approach,
same EBTEL modeling to the main flaring loops and no matter how we change the starting guesses for the
making a direct comparison of the heating rate between parameters, the synthetic peak in AIA 94 Å based on
the two sets of loops. Nevertheless, when comparing our the “best-fit” parameters is still systematically above
results with those revealed in other flares of the similar the observed peak, although the discrepancy between
16 Dai et al.

them is somewhat reconciled, and more importantly, evolution of the flare, in which different reconnections
the synthetic peak in AIA 335 Å is now systematically are involved and take place in different places. Using
lower than the observed peak (not shown here). Since the EBTEL model, we modeled the EUV emissions from
the light curve in neither AIA 335 Å nor AIA 94 Å can a late-phase loop. The modeling reveals a high heating
be reasonably modeled, the even worse fitting results rate for the late-phase loop. Our main conclusion is that
indicate that the late-phase emissions observed in AIA the extremely large late phase is powered by an intense
94 Å (and also in AIA 131 Å) are indeed affected by the heating even earlier than the main flare heating, and the
effect of non-equilibrium ionization. This also explains delayed occurrence of the late-phase peak is mainly due
why an EUV late phase as large as that in AIA 335 Å is to the long cooling process of the long late-phase loops.
not observed in a hotter passband, e.g., AIA 94 Å, even The production of EUV late phase in this flare is dif-
though the late-phase peak is clearly separated from the ferent from that in the 2011 September 6 X2.1 eruptive
main flare peak (Figure 1(d)). flare not only in timing of the heating process but also
In the cooler AIA passbands, the higher mass draining in the heating rate. In addition to these two flares, AR
rate during the radiative cooling stage can also cause the NOAA 11283 has produced a series of EUV late-phase
late-phase peak less prominent (see Figure 1(f)). On the flares, which exhibit different evolution patterns in the
other hand, we find that the synthetic intensities of the late-phase emissions. A comparative study of the pro-
late-phase loop in AIA 171 Å are systematically lower duction of the late phase in these flares will be presented
than the observed intensities. We propose that the dis- in a following work.
crepancy may be caused by a semi-circular assumption
of the loop geometry in EBTEL, which is not appropri-
We are very grateful of the anonymous referee whose
ate for the real geometry of the late-phase loop modeled valuable comments and suggestions led to a significant
here.
improvement of the manuscript. This work was sup-
ported by National Natural Science Foundation of China
6. SUMMARY under grants 11533005 and 11733003, and 973 Project
We have analyzed and modeled an M1.2 non-eruptive of China under grant 2014CB744203. D.Y. is also spon-
solar flare on 2011 September 9 that exhibits an ex- sored by the Open Research Project of National Center
tremely large EUV late phase in the warm coronal emis- for Space Weather, China Meteorological Administra-
sions. Based on an NLFFF extrapolation, we proposed a tion. SDO is a mission of NASA’s Living With a Star
two-stage magnetic reconnection scenario to explain the (LWS) Program.

REFERENCES
Antiochos, S. K., & Sturrock, P. A. 1976, SoPh, 49, 359 Cargill, P. J., Bradshaw, S. J., & Klimchuk, J. A. 2012,
—. 1978, ApJ, 220, 1137 ApJ, 752, 161
Asai, A., Yokoyama, T., Shimojo, M., et al. 2004, ApJ, 611, Cargill, P. J., Mariska, J. T., & Antiochos, S. K. 1995, ApJ,
557 439, 1034
Aschwanden, M. J., Nitta, N. V., Wuelser, J.-P., & Lemen, Carmichael, H. 1964, NASA Special Publication, 50, 451
J. R. 2008, ApJ, 680, 1477 Chamberlin, P. C., Milligan, R. O., & Woods, T. N. 2012,
Aulanier, G., Pariat, E., Démoulin, P., & DeVore, C. R. SoPh, 279, 23
2006, SoPh, 238, 347 Chen, P. F., Su, J. T., Guo, Y., & Deng, Y. Y. 2012,
Aulanier, G., Golub, L., DeLuca, E. E., et al. 2007, Science, Chinese Science Bulletin, 57, 1393
318, 1588 Dai, Y., & Ding, M. 2018, ApJ, 857, 99
Barnes, W. T., Cargill, P. J., & Bradshaw, S. J. 2016, ApJ, Dai, Y., Ding, M. D., & Guo, Y. 2013, ApJL, 773, L21
829, 31 Démoulin, P., Hénoux, J. C., Priest, E. R., & Mandrini,
Bobra, M. G., Sun, X., Hoeksema, J. T., et al. 2014, SoPh, C. H. 1996, A&A, 308, 643
289, 3549 Feng, L., Wiegelmann, T., Su, Y., et al. 2013, ApJ, 765, 37
Boerner, P., Edwards, C., Lemen, J., et al. 2012, SoPh, 275, Freeland, S. L., & Handy, B. N. 1998, SoPh, 182, 497
41 Galsgaard, K., & Nordlund, Å. 1997, J. Geophys. Res., 102,
Bradshaw, S. J., & Klimchuk, J. A. 2011, ApJS, 194, 26 219
Bradshaw, S. J., & Mason, H. E. 2003, A&A, 401, 699 Gary, G. A., & Hagyard, M. J. 1990, SoPh, 126, 21
Brueckner, G. E., Howard, R. A., Koomen, M. J., et al. Guo, Y., Ding, M. D., Schmieder, B., Démoulin, P., & Li,
1995, SoPh, 162, 357 H. 2012, ApJ, 746, 17
Calabretta, M. R., & Greisen, E. W. 2002, A&A, 395, 1077 Hirayama, T. 1974, SoPh, 34, 323
EUV LATE PHASE IN A NON-ERUPTIVE FLARE 17

Hock, R. A., Woods, T. N., Klimchuk, J. A., Eparvier, O’Dwyer, B., Del Zanna, G., Mason, H. E., Weber, M. A.,
F. G., & Jones, A. R. 2012, ArXiv e-prints, & Tripathi, D. 2010, A&A, 521, A21
arXiv:1202.4819 Pariat, E., & Démoulin, P. 2012, A&A, 541, A78
Hoeksema, J. T., Liu, Y., Hayashi, K., et al. 2014, SoPh, Parker, E. N. 1963, ApJS, 8, 177
289, 3483 Pesnell, W. D., Thompson, B. J., & Chamberlin, P. C.
Jiang, C., Wu, S. T., Feng, X., & Hu, Q. 2014, ApJ, 780, 55 2012, SoPh, 275, 3
Kaiser, M. L., Kucera, T. A., Davila, J. M., et al. 2008, Priest, E. R., & Longcope, D. W. 2017, SoPh, 292, 25
SSRv, 136, 5 Qiu, J., Liu, W.-J., & Longcope, D. W. 2012, ApJ, 752, 124
Kane, S. R., & Donnelly, R. F. 1971, ApJ, 164, 151 Qiu, J., Longcope, D. W., Cassak, P. A., & Priest, E. R.
Klimchuk, J. A., Patsourakos, S., & Cargill, P. J. 2008, 2017, ApJ, 838, 17
ApJ, 682, 1351 Reale, F., & Orlando, S. 2008, ApJ, 684, 715
Kopp, R. A., & Pneuman, G. W. 1976, SoPh, 50, 85 Romano, P., Zuccarello, F., Guglielmino, S. L., et al. 2015,
Lau, Y.-T., & Finn, J. M. 1990, ApJ, 350, 672 A&A, 582, A55
Leka, K. D., Barnes, G., Crouch, A. D., et al. 2009, SoPh, Ruan, G., Chen, Y., & Wang, H. 2015, ApJ, 812, 120
260, 83 Scherrer, P. H., Schou, J., Bush, R. I., et al. 2012, SoPh,
Lemen, J. R., Title, A. M., Akin, D. J., et al. 2012, SoPh, 275, 207
275, 17 Sturrock, P. A. 1966, Nature, 211, 695
Li, Y., Ding, M. D., Guo, Y., & Dai, Y. 2014, ApJ, 793, 85 Sun, X., Hoeksema, J. T., Liu, Y., et al. 2013, ApJ, 778, 139
Li, Y., Qiu, J., & Ding, M. D. 2012, ApJ, 758, 52 Titov, V. S., Hornig, G., & Démoulin, P. 2002, Journal of
Lin, R. P., Dennis, B. R., Hurford, G. J., et al. 2002, SoPh, Geophysical Research (Space Physics), 107, 1164
210, 3 Vemareddy, P., & Wiegelmann, T. 2014, ApJ, 792, 40
Liu, C., Deng, N., Lee, J., et al. 2014, ApJ, 795, 128 Wang, Y., Zhou, Z., Zhang, J., et al. 2016, ApJS, 223, 4
Liu, K., Wang, Y., Zhang, J., et al. 2015, ApJ, 802, 35 Wheatland, M. S., Sturrock, P. A., & Roumeliotis, G. 2000,
Liu, K., Zhang, J., Wang, Y., & Cheng, X. 2013, ApJ, 768, ApJ, 540, 1150
150 Wiegelmann, T. 2004, SoPh, 219, 87
Markwardt, C. B. 2009, in Astronomical Society of the Wiegelmann, T., Inhester, B., & Sakurai, T. 2006, SoPh,
Pacific Conference Series, Vol. 411, Astronomical Data
233, 215
Analysis Software and Systems XVIII, ed. D. A.
Woods, T. N. 2014, SoPh, 289, 3391
Bohlender, D. Durand, & P. Dowler, 251
Woods, T. N., Hock, R., Eparvier, F., et al. 2011, ApJ, 739,
Martens, P. C. H. 2010, ApJ, 714, 1290
59
Masson, S., Pariat, E., Aulanier, G., & Schrijver, C. J.
Woods, T. N., Eparvier, F. G., Hock, R., et al. 2012, SoPh,
2009, ApJ, 700, 559
275, 115
Masson, S., Pariat, É., Valori, G., et al. 2017, A&A, 604,
Wuelser, J.-P., Lemen, J. R., Tarbell, T. D., et al. 2004, in
A76
Proc. SPIE, Vol. 5171, Telescopes and Instrumentation
Metcalf, T. R., Leka, K. D., Barnes, G., et al. 2006, SoPh,
for Solar Astrophysics, ed. S. Fineschi & M. A. Gummin,
237, 267
111–122
Miklenic, C. H., Veronig, A. M., Vršnak, B., & Hanslmeier,
Xu, Y., Jing, J., Wang, S., & Wang, H. 2014, ApJ, 787, 7
A. 2007, A&A, 461, 697
Yang, K., Guo, Y., & Ding, M. D. 2015, ApJ, 806, 171
Milligan, R. O., Chamberlin, P. C., Hudson, H. S., et al.
Zhang, Q. M., Ning, Z. J., Guo, Y., et al. 2015, ApJ, 805, 4
2012, ApJL, 748, L14
Zhu, C., Qiu, J., & Longcope, D. W. 2018, ApJ, 856, 27
Neupert, W. M. 1968, ApJL, 153, L59

You might also like